Download Unearthing the Roles of Imprinted Genes in the Placenta

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Non-coding DNA wikipedia , lookup

Genomic library wikipedia , lookup

Point mutation wikipedia , lookup

Human genome wikipedia , lookup

Genomics wikipedia , lookup

Transgenerational epigenetic inheritance wikipedia , lookup

Epigenomics wikipedia , lookup

Gene therapy of the human retina wikipedia , lookup

Gene nomenclature wikipedia , lookup

Public health genomics wikipedia , lookup

Pathogenomics wikipedia , lookup

Gene therapy wikipedia , lookup

Gene desert wikipedia , lookup

Genetic engineering wikipedia , lookup

Biology and consumer behaviour wikipedia , lookup

Fetal origins hypothesis wikipedia , lookup

Ridge (biology) wikipedia , lookup

Oncogenomics wikipedia , lookup

Behavioral epigenetics wikipedia , lookup

Epigenetics wikipedia , lookup

Cancer epigenetics wikipedia , lookup

Vectors in gene therapy wikipedia , lookup

Minimal genome wikipedia , lookup

Cell-free fetal DNA wikipedia , lookup

Long non-coding RNA wikipedia , lookup

Gene wikipedia , lookup

RNA-Seq wikipedia , lookup

X-inactivation wikipedia , lookup

Gene expression programming wikipedia , lookup

NEDD9 wikipedia , lookup

Therapeutic gene modulation wikipedia , lookup

Epigenetics of diabetes Type 2 wikipedia , lookup

Epigenetics in learning and memory wikipedia , lookup

Genome evolution wikipedia , lookup

Epigenetics in stem-cell differentiation wikipedia , lookup

Polycomb Group Proteins and Cancer wikipedia , lookup

Epigenetics of neurodegenerative diseases wikipedia , lookup

History of genetic engineering wikipedia , lookup

Genome (book) wikipedia , lookup

Microevolution wikipedia , lookup

Gene expression profiling wikipedia , lookup

Artificial gene synthesis wikipedia , lookup

Designer baby wikipedia , lookup

Epigenetics of human development wikipedia , lookup

Site-specific recombinase technology wikipedia , lookup

Nutriepigenomics wikipedia , lookup

Genomic imprinting wikipedia , lookup

Transcript
Placenta 30 (2009) 823–834
Contents lists available at ScienceDirect
Placenta
journal homepage: www.elsevier.com/locate/placenta
Current Topic
Unearthing the Roles of Imprinted Genes in the Placenta
F.F. Bressan a, T.H.C. De Bem a, F. Perecin a, F.L. Lopes b, C.E. Ambrosio a, F.V. Meirelles a, M.A. Miglino c, *
a
Department of Basic Sciences, Faculty of Animal Sciences and Food Engineering, University of São Paulo, Pirassununga, Brazil
Department of Human Genetics, McGill University, Montreal, Canada
c
Department of Surgery, Faculty of Veterinary Medicine and Animal Sciences, University of São Paulo, São Paulo, Brazil
b
a r t i c l e i n f o
a b s t r a c t
Article history:
Accepted 22 July 2009
Mammalian fetal survival and growth are dependent on a well-established and functional placenta.
Although transient, the placenta is the first organ to be formed during pregnancy and is responsible for
important functions during development, such as the control of metabolism and fetal nutrition, gas and
metabolite exchange, and endocrine control. Epigenetic marks and gene expression patterns in early
development play an essential role in embryo and fetal development. Specifically, the epigenetic
phenomenon known as genomic imprinting, represented by the non-equivalence of the paternal and
maternal genome, may be one of the most important regulatory pathways involved in the development
and function of the placenta in eutherian mammals. A lack of pattern or an imprecise pattern of genomic
imprinting can lead to either embryonic losses or a disruption in fetal and placental development.
Genetically modified animals present a powerful approach for revealing the interplay between gene
expression and placental function in vivo and allow a single gene disruption to be analyzed, particularly
focusing on its role in placenta function. In this paper, we review the recent transgenic strategies that
have been successfully created in order to provide a better understanding of the epigenetic patterns of
the placenta, with a special focus on imprinted genes. We summarize a number of phenotypes derived
from the genetic manipulation of imprinted genes and other epigenetic modulators in an attempt to
demonstrate that gene-targeting studies have contributed considerably to the knowledge of placentation
and conceptus development.
Ó 2009 Elsevier Ltd. All rights reserved.
Keywords:
Epigenetics
Genomic imprinting
Knockout
Placentation
Transgenesis
1. Introduction
In mammals, embryo development and survival, as well as
a successful pregnancy, are dependent on the establishment of
a functional maternal–fetal interface. This connection is initiated
during the primary contact of the embryo, followed by embryo
implantation, which is characterized by fetal trophoblast cell
invasion into the maternal endometrium, and it culminates with
the generation of the chorioallantoic placenta [reviewed by [1]].
Together, these processes are referred to as placentation [2].
The phenomenon of genomic imprinting has been demonstrated extensively to play a key role in fetal development and
placentation [3,4]. Although the majority of imprinted genes are
expressed in extraembryonic tissues, there is little information
available on the mechanisms by which such mono-allelic gene
expression regulates placental growth, development and function
[5,6].
Continuous research on placentation and the myriad mechanisms controlling this process is needed to clarify the embryonic–
endometrial interactions, and the use of animal models has
contributed greatly to this study [7]. In particular, genetically
modified animals have provided much of the knowledge on the
genetic control of placental development [8]. In fact, the use of
transgenic models has enabled the creation and analysis of gene
regulation assays; the discovery of new roles for genes in placentation; and, most importantly, it has contributed to our understanding of developmental and perinatal pathologies in animals
and humans. In the present review, we address the epigenetic
events involved in embryogenesis, focusing on imprinted genes
and the knowledge generated by transgenic models as tools to
increase our understanding of the roles that imprinted genes play
in placentation and early development.
2. Epigenetics and development
* Corresponding author. Tel.: þ55 11 30917690.
E-mail address: [email protected] (M.A. Miglino).
0143-4004/$ – see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.placenta.2009.07.007
The placenta is the first organ to be formed during pregnancy. It
is responsible for the establishment of vascular connections
between mother and conceptus and allows for the exchange of gas,
824
F.F. Bressan et al. / Placenta 30 (2009) 823–834
nutrients and waste. This organ is involved in immune protection of
the fetus and also produces the hormones needed to support fetal
development [9].
The creation of an appropriate maternal environment for fetal
development depends on the proper functioning and development
of the trophoblast cells, which require the well-coordinated
expression of many transcription factors, cell cycle regulators,
growth factors, cytokines and surface receptors [reviewed by
[10,11]]. Embryogenesis and placentation are particularly prone to
perturbations in gene expression because these processes depend
on a complex cascade of events [12,13]. Any disruption to the wellorchestrated expression of these regulatory factors may lead to
placental disorders, causing undesirable phenotypes or even
precocious deaths in animals or humans [9].
Following fertilization, a single-cell zygote forms a multicellular
organism comprised of more than 200 different cell types [14,15].
The development of lineage-specific cells begins with the differentiation of the trophoblast lineage and the inner cell mass [16].
This event depends on epigenetic modifications that control the
expression of particular genes, allowing cells to develop and
differentiate into specific cells and tissues [17].
Epigenetics can be defined as the heritable changes in gene
expression that are not caused by the changes in DNA sequence
[18]. The best studied epigenetic mechanisms are DNA methylation
and histone post-translational modifications, which interact with
each other and also with regulatory proteins and non-coding RNAs
[reviewed by [19]].
The paternal genome is actively demethylated within a few
hours of fertilization, while the maternal genome is demethylated
passively during the first cleavages in a species-dependent manner.
This demethylation, however, spares imprinted genes [20], which
must be maintained throughout development without being
‘‘de novo’’ reprogrammed during the pre-implantation stages [21].
Imprinted genes are expressed selectively from either the
paternal or maternal allele. This specialized form of gene regulation
is necessary for normal development [22,23], as discussed below. In
paternally imprinted genes, the paternal allele is epigenetically
modified, preventing its transcription and leading to mono-allelic
maternal expression [18,24]. The same happens to the maternally
imprinted genes, in which the paternal allele is solely expressed.
These selectively expressed genes are believed to have an important role in the allocation of maternal resources to fetal growth
[25,26].
Imprinted genes are found throughout the mammalian genome,
though their occurrence is not random. These genes tend to be
found in clusters that contain DNA sequences that are rich in CpG
nucleotides. These specific regions, called imprinting control
regions (ICRs), are characterized by epigenetic marks, mainly DNA
methylation and histone modifications, which influence the
binding affinity of transcription activators/suppressors and recruit
chromatin remodeling enzymes to locally change the structure and
function of chromatin [27]. The existence of control regions
suggests that genomic imprinting may be controlled not only at the
single gene level but at the level of the chromosome [28].
Epigenetic marks present in single parental copies of imprinted
regions are responsible for differential gene expression. Interestingly, the maintenance of imprinting has been recently inferred to
depend more on repressive histone methylation than on DNA
methylation in the placenta [6,29].
3. Genomic imprinting and placental development
Approximately 200 genes are imprinted in the mammalian
genome [30]. More than 70 imprinted genes in mice and at least 50 in
humans have already been reported in the current literature (http://
www.mgu.har.mrc.ac.uk/imprinting, http://www.geneimprint.com,
http://igc.otago.ac.nz). In most genes, the imprinting status is
conserved between mouse and human [25] and in some genes the
imprinted status is reported to be conserved also in other species, i.e.,
cattle [31–34]. As summarized in Table 1, imprinted gene expression
can be found in the placenta, the fetus, or both, independently of the
parental origin of the expressed allele, and may be widespread or
specific to certain cell types [4]. Although imprinted gene functions
are generally essential for the proper development and function of
the placenta, as well as for fetal growth [6], some of these genes have
not been reported to be related to development. It is important to
note, however, that imprinted genes can show spatial-temporal
expression [35]. Their expression window during development,
therefore, may be narrow enough to cause the imprinted characteristic to be difficult to recognize.
The placenta is one of the most important sites of imprinted
gene action [[36] reviewed by [37]]. Although placentation displays
species-specific variation [2], the genomic imprinting phenomenon
is conserved amongst eutherian mammals, especially primates,
rodents and ruminants [6,38].
According to the conflict hypothesis [39,40], paternally
expressed genes enhance fetal growth, while maternally expressed
genes suppress fetal growth. One evolutionary explanation for this
hypothesis would be that by restricting fetal growth, females can
have a longer reproductive lifespan, assuring their reproductive
success. In contrast, having more numerous and stronger progeny is
advantageous for males. The conflict hypothesis achieved some
confirmation through observations made with mouse genome
manipulation. Androgenote mice, which contain only paternal
DNA, have poorly developed embryonic components but better
developed extraembryonic tissues, whereas gynogenotes show the
opposite phenotype [41].
It is important to note that both the accurate establishment of
genomic imprints and the correct maintenance of genomic
imprints during embryogenesis are essential for normal embryonic/placental development [42]. Epimutations affecting
imprints can arise during imprint erasure, which occurs when
germ cells migrate to the gonads in pre-natal stages, during
either the imprint establishment that takes place during gametogenesis or imprint maintenance throughout the life of the
organism [43,44].
A clear example of epigenetic disturbance in development is the
interference caused by assisted reproductive techniques (ARTs).
These techniques likely interfere with imprint establishment
(manipulation of gametes) or imprint maintenance (manipulation
of pre-implantation embryos; [43]).
4. Imprinted genes control mammalian development
Insulin-like growth factor 2 (Igf2) was one of the first imprinted
genes to be discovered [45]. Igf2 and its receptor, Igf2r, are essential
during fetal–placental development [46]. While the former is
a maternally imprinted gene that codes for a growth factor involved
in fetal and placental growth in mice and humans, the latter is
a maternally expressed gene in mice involved in Igf2 degradation.
Although recent studies demonstrated that IGF2r is not imprinted
in humans [47,48], the relationship between these genes brings
strength to the conflict theory [49,50].
Igf2, together with H19, which is an imprinted non-coding
transcript, is located in a cluster of imprinted genes in mouse
chromosome 7, syntenic to human chromosome 11p15.5 [51,52]. A
region upstream of H19 regulates imprinted expression of both of
these genes [53]. The establishment and maintenance of DNA
methylation in the Igf2/H19 DMR is acquired during spermatogenesis in the male germ cells; however, the DMR from the female
F.F. Bressan et al. / Placenta 30 (2009) 823–834
825
Table 1
Imprinted gene expression reported in mouse development.
Gene
Aliases
Chromosome
location
Preferentially
imprinted allele
Name
References
Gatm
Nnat
Nesp
AT
Peg 5
Central 2
Distal 2
Distal 2
Paternal
Maternal
Paternal
L-Arginine:glycine amidinotransferase
Neuronatin
Neuroendocrine secretory protein
Distal 2
Maternal
Distal 2
Maternal
Neuro endocrine secretory
protein antisense
Guanine nucleotide binding protein,
alpha stimulating
[140] (Extraembryonic tissues)
[141] (Fetal brain)
[142,143] (Embryonic and
extraembryonic tissues)
[143,144] (embryonic tissues)
Distal 2
Maternal
Distal 2
Distal 2
Maternal
Paternal
Guanine nucleotide binding protein,
alpha stimulating, ‘extra large’
Malignant T-cell amplified sequence 2
Histocompatibility 13
Proximal 2
Maternal
Scm-like with four mbt domains 2
Nespas
Gnas
Gs-alpha
Gnasxl
Mcts2
H13
SPP
Sfmbt2
Calcr
Mit1/Lb9
Clr
Proximal 6
Proximal 6
Paternal
Maternal
Calcitonin receptor
Mest-linked imprinted transcript 1
Sgce
e-SG
Proximal 6
Maternal
Sarcoglycan, epsilon
Peg10
Edr, HB-1, Mar2,
MEF3L, Mart2, MyEF-3
Proximal 6
Maternal
Paternally expressed gene 10
Ppp1r9a
Pon3
Pon2
Asb4
Proximal
Proximal
Proximal
Proximal
Paternal
Paternal
Paternal
Paternal
Mest/Peg1
Proximal 6
Maternal
Neurabin
Paraoxonase 3
Paraoxonase 2
Ankyrin repeat and suppressor
of cytokine signalling
Mesoderm specific transcript
Copg2
Proximal 6
Paternal
Proximal 6
Proximal 6
Maternal
Paternal
Coatomer protein complex subunit
gamma 2
Copg2 antisense
Kruppel-like factor 14
Kvlqt1as
Proximal 6
Proximal 7
Proximal 7
Distal 7
Maternal
Maternal
Paternal
Maternal
Nucleosome assembly protein 1-like 5
Zinc-finger gene 264
Zinc-finger gene 3 from imprinted domain
Kvlqt1 antisense
Pw1, End4, Gcap4, Zfp102
Ocat
Proximal
Proximal
Proximal
Proximal
Paternal
Paternal
Maternal
Maternal
Imprinted zinc-finger gene 3
Imprinted zinc-finger gene 1
Paternally expressed gene 3, probably Pw1
Ubiquitin-specific processing protease 29
Copg2as
Klf4
Nap1l5
Zfp264
Zim3
Kcnq1ot1
Zim2
Zim1
Peg3
Usp29
Ube3a
Epfn, Klf14, epiprofin, BTEB5
Znf264
6
6
6
6
7
7
7
7
Hpve6a, E6-AP ubiquitin
protein ligase
snoRNA MBII-85, Snord116
Peg4, HCERN3
Central 7
Paternal
E6-Ap ubiquitin protein ligase 3A
Central 7
Central 7
Maternal
Maternal
Peg6
ns7, nM15, NDNL1, Mage-l2
Central
Central
Central
Central
7
7
7
7
Maternal
Maternal
Maternal
Maternal
Prader–Willi chromosome region 1
Small nuclear ribonucleoprotein
polypeptide N (Snrpn),
Snrpn upstream reading frame (Snurf)
Paternally expressed in the CNS 2
Paternally expressed in the CNS 3
necdin
Melanoma antigen, family L, 2
Central 7
Central 7
Maternal
Maternal
Peg12/Frat3
Central 7
Maternal
Inpp5f_v2
Distal 7
Maternal
Inpp5f_v3
Distal 7
Maternal
H19
Distal 7
Paternal
Pwcr1
Snrpn/Snurf
Pec2
Pec3
Ndn
Magel2
Mkrn3
Zfp127as/Mkrnas
Zfp127
Ring zinc-finger encoding gene 127
Ring zinc-finger encoding gene
127 antisense
Frequently rearranged in advanced
T-cell lymphomas
Inositol polyphosphate-5-phosphatase,
variant 2
Inositol polyphosphate-5-phosphatase,
variant 3
[145] (Embryonic tissues,
predicted by the embryonic lethality
of null mutations)
[142,146] (Embryonic tissues)
[147] (Embryonic tissues)
[147] (Embryonic and
extraembryonic tissues)
[148] (Early embryos and
extraembryonic tissues)
[149](Fetal brain)
[150] (Fetal brain, partially
imprinted in other fetal tissues)
[65,151] (Embryonic and
extraembryonic tissues)
[65] (Embryonic and
extraembryonic tissues)
[65,152] (Extraembryonic tissues)
[65,152] (Extraembryonic tissues)
[65,152] (Placenta-specific)
[153] (Embryonic and
extraembryonic tissues)
[154,155] (Embryonic and
extraembryonic tissues)
[150] (Embryonic tissues)
[150] (Embryonic tissues)
[156] (Embryonic and
extraembryonic tissues)
[157] (Embryonic tissues)
[158] (Embryonic tissues)
[158] (Embryonic tissues)
[134,159] (Embryonic and
extraembryonic tissues)
[160] (Embryonic tissues)
[161] (Embryonic tissues)
[161,162] (Embryonic tissues)
[163,164] (Mid-gestation
embryos, fetal brain)
[164] (Fetal brain)
[165] (Embryonic tissues)
[166–168] (Embryonic and
extraembryonic tissues)
[164] (Fetal brain)
[164] (Fetal brain)
[164,169] (Fetal brain)
[170] (Extraembryonic tissues
and fetal brain)
[171,172] (Embryonic tissues)
[173] (Pre-implantation embryo)
[174] (Embryonic tissues)
[175] (Fetal brain)
[147] (Fetal brain)
Igf2
Mpr, M6pr, Peg2, Igf-2, Igf-II
Distal 7
Maternal
Insulin-like growth factor type 2
Ins2
Mody, Ins-2, InsII, Mody4,
proinsulin, INS
Distal 7
Maternal
Insulin 2
[176,177] (Embryonic and
extraembryonic tissues)
[45] (Embryonic and
extraembryonic tissues)
[178,179] (Extraembryonic tissues)
Distal 7
Paternal
Mus musculus achaete-scute homologue 2
[135] (Placenta-specific)
Ascl2/Mash2
(continued on next page)
826
F.F. Bressan et al. / Placenta 30 (2009) 823–834
Table 1 (continued )
Gene
Aliases
Chromosome
location
Preferentially
imprinted allele
Name
References
Tapa1/Cd81
Tssc4
Tspan28
Distal 7
Distal 7
Paternal
Paternal
[133] (Extraembryonic tissues)
[134,180] (Placenta-specific)
Kcnq1
Kvlqt1
Distal 7
Paternal
Cdkn1c
p57Kip2
Distal 7
Paternal
cd 81 antigen
Tumor-suppressing subchromosomal
transferable fragment 4
Potassium voltage-gated channel,
subfamily Q, member 1
Cyclin-dependent kinase inhibitor 1C
Slc22a18
Distal 7
Paternal
Solute carrier family 22, member 18
Phlda2
HET, ITM, Impt1, TSSC5, Orctl2,
Slc22a1l, Slc22a1, BWR1A
Ipl, Tssc3
Distal 7
Paternal
Nap1l4
Nap2
Distal 7
Paternal
Pleckstrin homology-like domain,
family A, member 2 (Phlda2),
Imprinted in placenta and liver (Ipl)
Nucleosome assembly protein 1-like 4
Tnfrsf23
Tnfrh1
Distal 7
Maternal
Obph1
Osbpl5
Distal 7
Paternal
Plagl1
Dcn
Lot1, Zac1
DC, PG40, PGII,
PGS2, mDcn,
DSPG2, SLRR1B
Aadc
Proximal 10
Central 10
Maternal
Paternal
Proximal 11
Maternal
Meg 1
SP2, 35 kDa, Irlgs2,
D11Ncvs75,
U2afbp-rs, Zrsr1
Meg9
Proximal 11
Proximal 11
Paternal
Maternal
Distal 12
Paternal
Dopa decarboxylase (Ddc); aromatic
L-amino acid decarboxylase (Aadc)
Growth factor receptor bound protein
U2 small nuclear ribonucleoprotein
auxiliary factor (U2AF), 35 kDa,
related sequence 1
miRNA containing gene
FA1, ZOG, pG2,
Peg9, SCP1,
Ly107, pref-1
Meg 3
Distal 12
Maternal
Delta-like 1
Distal 12
Maternal
Gene trap locus 2
Mar, Mor1, Mart1,
Peg11
Distal 12
Maternal
Retrotransposon-like 1
Distal 12
Maternal
Deiodinase, iodothyronine type III
Ddc
Grb10
U2afl1-rs1
Mirg
Dlk1
Gtl2
Rtl1
Dio3
Tumor necrosis factor receptor
superfamily, member 23
Oxysterol-binding protein 1 (Obph1),
oxysterol binding
protein-like 5 (Osbpl5)
Pleomorphic adenoma gene-like 1
Decorin
Antipeg11/Rtl1as
Hosts several miRNAs
Distal 12
Paternal
Antisense to Rtl1/Peg11
Htr2a
Htr2, Htr-2, 5-HT2A
receptor
Task3
Distal 14
Paternal
Distal 15
Paternal
Distal 15
Maternal
5-Hydroxytryptamine
(serotonin) receptor 2 A
Potassium channel, subfamily K,
member 9
Paternally expressed 13
Kcnk9
Peg13
Slc238a4
Ata3, mATA3
Distal 15
Maternal
Slc22a3
EMT, Oct3, Orct3, Slca22a3
Proximal 17
Paternal
Slc22a2
Oct2, Orct2
Proximal 17
Paternal
Igf2r
CD222, CI-MPR, Mpr300,
M6P/IGF2R
Air, Igf2ras
Proximal 17
Paternal
Proximal 17
Maternal
Airn
germline cell is protected against methylation by the zinc-finger
protein CTCF [52]. Such protection prevents interactions between
the Igf2 gene and enhancers located downstream of H19 in the
maternal allele, thus preventing Igf2 transcription. When CTCF does
not bind to the paternal allele, on the other hand, Igf2 is expressed,
and DNA is methylated within the H19 promoter region, resulting in
H19 transcriptional silencing. The different methylation status of
the Igf2–H19 locus, therefore, guarantees the exclusive paternal Igf2
expression and maternal H19 expression [51].
The importance of the parental origin of Igf2/H19 genes was
elegantly demonstrated when Kono and collaborators (2004, [54])
Solute carrier family 38,
member 4/amino
acid transport system A3
Solute carrier family 22
(organic cation transporter), member 3
Solute carrier family 22
(organic cation transporter), member 2
Insulin-like growth factor type 2 receptor
Insulin-like growth factor 2
receptor antisense RNA
[134,180,181] (Embryonic and
extraembryonic tissues)
[135,182] (Embryonic and
extraembryonic tissues)
[183,184] (Embryonic and
extraembryonic tissues)
[185,186] (Weakly in embryonic,
mainly in extraembryonic tissues)
[181] (Mainly in placenta; however,
reported not imprinted by [187])
[188] (Embryonic and
extraembryonic tissues)
[181,189] (Placenta-specific)
[151] (Embryonic tissues)
[153] (Placenta)
[190] (Embryonic heart)
[191] (Embryonic tissues)
[192] (Embryonic tissues)
[193] (Embryonic and
extraembryonic tissues)
[194,195] (Embryonic and
extraembryonic tissues)
[195,196] (Embryonic and
extraembryonic tissues)
[196] (Embryonic and
extraembryonic tissues)
[197] (Embryonic tissues and
weakly imprinted in
extraembryonic tissues)
[198] (Embryonic and
extraembryonic tissues)
[199] (Embryonic eye)
[200] (Embryonic tissues)
[157] (Embryonic and
extraembryonic tissues)
[153] (Embryonic and
extraembryonic tissues)
[201] (Placenta-specific)
[202] (Placenta-specific)
[203,204] (Embryonic and
extraembryonic tissues)
[57,205] (Embryonic and
extraembryonic tissues)
successfully produced viable parthenogenetic offspring in mice by
correcting the Igf2/H19 dosage. In this experiment, one of the
maternal alleles was derived from a non-growing oocyte (ng),
while the other was derived from a fully grown (fg) oocyte. The
process of imprinting in the maternal germline occurs at late stages
of oogenesis. Therefore, ng oocytes are considered to be ‘‘imprintneutral’’, and both H19 and Igf2 genes are expressed [43,55,56]. By
introducing a deletion in the H19 gene and its flanking regions in
the ng oocyte and consequently disrupting the imprinting of Igf2
gene, the authors demonstrated both that parthenogenetic development to term could be achieved and also that the proper
F.F. Bressan et al. / Placenta 30 (2009) 823–834
expression of Igf2/H19 likely drove modifications of other genes
that allow parthenote survival.
The Igf2r cluster, which contains Slc22a2 and Slc22a3 genes,
a solute carrier family 22 that codifies imprinted genes, is also
regulated by methylation-sensitive elements. Unlike most imprinted genes, the methylated allele is expressed in this cluster. In this
gene, the maternally methylated allele leads to paternal Igf2r
repression. The paternal non-methylated allele expresses a noncoding RNA (ncRNA), called Airn (previously named Air), which is
responsible for preventing paternal Igf2r expression [52,57].
Other important imprinted loci display the same behavior. The
Gnas and Kcnq1 loci, for example, contain ncRNAs believed to
contribute to genomic imprint control, i.e., Nespas/Gnas-as and
Kcnq1ot1, respectively. Therefore, in addition to DNA methylation
and post-translational histone modification, ncRNAs also control
imprinted gene expression [58].
The mechanisms by which ncRNAs are responsible for the
epigenetic changes observed in these imprinted loci are still not
well characterized. Numerous ncRNAs are located in clusters
regulated by ICRs [59]. In fact, each imprinted region expresses at
least one ncRNA [58,60]. Although their function and mechanisms
are not well understood, it is known that ncRNAs regulate
imprinted clusters that recruit chromatin remodeling complexes to
nearby genomic regions. The expression of specific ncRNAs, i.e.,
long ncRNAs, is associated with the acquisition of genomic
imprinting and the silencing of imprinting clusters [61,62].
827
A recently discovered imprinted retrotransposon-derived gene,
Peg10 [63], showed an essential function as an endogenous gene in
placental development [64]. Peg10 is highly conserved among
mammalian species [65], raising questions about its importance in
mammalian evolution. Ono and collaborators [64] highlighted the
possibility that ancestral mammals may have developed placenta
from newly acquired retrotransposon-derived genes or by modification of endogenous genes present in oviparous animals millions
of years ago. The understanding of the physiological roles of Peg10
and the other imprinted retrotransposon homologue Rtl1 is definitely important to improving our understanding of placental
evolution.
Disrupting the normal regulation of imprinted genes is decisive
throughout gestation and post-natal life, often leading to lethal
phenotypes in early development, as described in Table 2. Not
surprisingly, these phenotypes are related to several human
syndromes and disorders in post-natal life.
The IGF2 gene, for example, is involved in Russell–Silver
syndrome (RSS), which is characterized by the loss of methylation
in IGF2–H19 ICR, reduction in IGF2 expression, and biallelic
expression of H19, resulting in intrauterine and post-natal growth
retardation [66]. Beckwith–Wiedemann syndrome (BWS), on the
other hand, is characterized by the loss of IGF2 imprinting, causing
biallelic overexpression and a lack of expression of H19, leading to
overgrowth of the fetus, among other symptoms. Both BWS and
RSS phenotypes include pronounced growth disorders [67].
Table 2
Imprinted genes knockout and their phenotypes.
Imprinted gene
Mouse KO phenotype
References
Nesp
Gnas
Development without any obvious phenotype – behavior linked
Embryonic lethality. Heterozygous disruption is associated with significant early post-natal lethality. When maternal allele is disrupted
mice become obese. When paternal allele is disrupted, mice are hypermetabolic and thin
Increased myoclonus and deficits in motor coordination and balance
Growth retardation and early embryonic lethality due to incomplete placenta formation
Reduction in contextual fear memory, loss of hippocampal long-term potentiation
Embryonic and placental growth retardation
Neonatal lethality within 15 hours of birth, selective perturbation of late-stage differentiation structures in the epidermis
Reduction of 10–20% of weight
Embryonic and placental growth retardation, impairment of normal maternal behavior
Motor dysfunction, inducible seizures, context-dependent learning deficit
Severe post-natal growth retardation, delayed sexual maturation, but fertile. Elevated level of anxiety or fear. Motor learning deficiency,
hyperphagie
Viable offspring, with no obvious phenotypic or histopathologic defects. However, KO of its IC leads to increase in neonatal mortality and
underweight newborns showing hypotonia
Neonatal lethality and respiratory distress, underweight at birth
Reduced viability at embryonic day 12.5. Offspring showing disregulation of sleep and food intake, growth retardation soon after birth
Viable, healthy and fertile. No obvious phenotype. Triple Frat knockout (Frat1, Frat2 and Frat3) shows the same normal phenotype
Increase in placental weight, fetal overgrowth
P0 and null mutants showed reduced placental growth, followed by fetal growth restriction. Phenotypes more severe in Igf2 null
mutants at later stages of gestation
Viable and fertile, without major metabolic disorders. Ins1 and Ins2 double homozygous knockout, however, were growth-retarded,
developed diabetes mellitus and died within 48 h
Death at 10 d post-coitum, placental failure
Reduction of female fertility, increase in post-natal lethality
Deafness, circular movement and repetitive falling. Gastric hyperplasia. Severe anatomic disruption of cochlear and vestibular end
organs. Phenotypes unrelated to BWS
Divergent phenotypes in offspring. Abnormal placental development (placentomegaly and trophoblast dysplasia), morphological
defects in neonates
Placental overgrowth, consequent reduction of fetal-to-placental weight ratio
Intrauterine growth restriction, altered bone formation, increased neonatal lethality
Skin fragility, tumor development
Embryo and placenta overgrowth
Pre- and post-natal growth retardation, eyelid and skeletal abnormalities, smaller litter size, increased neonatal mortality
Fetal and post-natal growth reduction
Placental abnormalities and functional deficiencies, pre- and post-natal growth retardation, placental growth retardation, increased
late-fetal or neonatal lethality
Placental and fetal growth restriction
Impairment of neurotransmitters release. No obvious phenotypes, viable and fertile offspring
No obvious phenotypes, viable and fertile offspring
Lethality at birth, embryo overgrowth
Reduction in birth weight
[206]
[145,207]
Sgce
Peg10
Ppp1r9a
Mest/Peg1
Klf4
Kcnq1ot1
Peg3
Ube3a
Pwcr1
Snrpn
Ndn
Magel2
Peg12/Frat3
H19
Igf2
Ins2
Ascl2/Mash2
Tapa1/Cd81
Kcnq1
Cdkn1c
Phlda2
Plagl1
Dcn
Grb10
Dlk1
Gtl2
Rtl1
Slc238a4
Slc22a3
Slc22a2
Igf2r
Air
[208]
[64]
[209]
[210]
[211]
[212]
[213]
[214]
[215]
[216]
[217,218]
[219]
[220]
[221]
[26,46]
[222,223]
[224]
[225]
[226]
[227–229]
[186]
[230]
[231,232]
[233]
[234]
[235]
[196]
[236]
[202]
[237]
[238]
[57,239]
828
F.F. Bressan et al. / Placenta 30 (2009) 823–834
Abnormal imprinting patterns are also associated with neurodevelopmental disorders, such as Prader–Willi (PWS) and Angelman (AS) syndromes, which are associated with the loss of paternal
or maternal imprinting on chromosome 15q11–q13, respectively
[reviewed by [14,23,68]].
context, the generation of in vivo gene function assays is vital for
understanding the biological roles of developmental genes and
their interactions with each other and with environmental stimuli.
5. Imprinting alterations and implications
Understanding the genetic control of fetal–maternal interactions has dramatically improved with the introduction of genome
modifications in animal models. In fact, gene-targeting strategies
are the most widely accepted models used to provide reliable and
accurate information on the mechanisms of implantation and
placentation, given their ability to provide definitive evidence for
the in vivo function of a specific gene.
Genes that are candidates to have a role in early development
can have their biological effects analyzed in vivo in one of the two
ways: gain of function or loss of function studies. The first method
is based on gene overexpression, achieved by the random integration of a transgene into the genome or a targeted insertion of the
transgene into a specific locus (a knock-in). On the contrary, the loss
of function gene assay relies on the suppression of a gene function.
Mainly, it is achieved by gene-trapping in ES cells or targeted gene
deletion (a knockout, KO). The first method, although relatively
inexpensive, has the significant limitation of being only effective for
genes that are expressed in ES cells, whereas gene targeting can be
used for any gene, either permanently or in a conditional manner
[reviewed by [94–96]].
The gain of function strategy is especially interesting for characterizing placental features that are not fully described. The
transfer of transgenic embryos expressing a reporter gene, such as
green fluorescent protein (GFP) or the b-galactosidase enzyme
(LacZ), to wild-type recipients enables the precise discrimination of
uterine and trophoblast contributions to placental defects [97]. The
inverse is also valid when wild-type blastocysts are introduced into
mutant uterine tract [71]. This technique has been used for several
purposes, such as elucidating trophoblast invasion in hemochorial
placentas [98], demonstrating the spatial-temporal pattern of
imprinted gene expression in embryos [99] or revealing the X
inactivation mechanism [100,101].
KO mice model is another strategy that has greatly contributed
to the understanding of several diseases and different biological
processes [reviewed by [94]], usually revealing a gene role by
comparing the knockout phenotype with that of wild-type mice.
For example, it has been used to uncover basic mechanisms of DNA
repair [102], cancer research [103], diabetes [104], behavioral
analysis [105], and developmental related processes [106,107],
among several others.
Despite differences between mice and human morphology and
endocrine function, the mouse is the most popular model organism
for studying mammalian genomic imprinting and other processes
in eutherian animals [108]. Great advantages of mice when
compared to other animals are the availability of maternal- or
paternal-only derived embryos and the characteristics of these
animals, such as uniparental chromosomal duplications (UPD),
high fertility, low costs to maintain feeding and housing facilities,
and responsiveness to a range of assisted reproductive technologies
[94,109]. Most importantly however, is the availability of a fully
sequenced genome for this species [110] and the technology
available for the manipulation of embryonic stem cells, allowing
the use of these cells for the production of genetically altered
offspring [111,112].
The generation of KO mice relies on several in vitro procedures
that, although specific, are technically simple to perform. The first
step consists of the design and construction of the desired vector.
Circular sections of bacterial DNA (plasmids) are frequently used to
manipulate the genome of embryonic stem cells by introducing
In humans, pregnancy losses are extremely common and not
completely understood. In fact, 25% of spontaneous abortions
remain unexplained [69]. The majority of these losses occur during
the pre-implantation period, though after implantation, approximately 15–20% of pregnancies are also lost spontaneously [70,71].
In farm animals, embryonic mortality is also the major cause of
reproductive wastage, where a dysfunctional placenta accounts for
80% of this mortality [72,73].
ARTs have been widely used in an attempt to correct fertility
impairment in humans and animals and to provide a higher
reproductive efficiency in farm animals. In 2003, almost 4% of the
total number of human births in developed countries was estimated to have been produced with in vitro procedures [74]. This
scenario is not different for farm species. The last report of the IETS
(International Embryo Transfer Society), released in 2006,
announced that in the previous year, nearly 266,000 bovine
embryos were produced in vitro and transferred worldwide.
Despite its wide use, ARTs, such as IVF or cloning in animals,
increase the incidence of abnormalities in the morphology and
function of the placenta [75]. Hydroallantois, poor vascularization
and abnormal (mostly reduced but also enlarged) placentomes are
some of the most common pathological alterations [76–78].
Overall growth of the placenta and other particular structures
(such as the labyrinthine trophoblast), as well as regulation of
specific transporters and channels needed for nutrient supply to
the fetus, are frequently regulated or affected by imprinted genes
[reviewed by [25]].
Placental perturbations also lead to high birth weights and
reduced survival rates, a condition known in ruminants as large
offspring syndrome (LOS, [79,80]). This condition is reminiscent of
the BWS in humans and is correlated with IGF2R imprinting
disruption [81]. The incidence of placental failures is especially
important in cloning by nuclear transfer because such failures
represent the major cause of pregnancy failure in these animals
[76,82–85]. Placental abnormalities in cloned animals are evident
and appear frequently even in gestations carried to term [86,87].
Furthermore, the use of ARTs and their in vitro culture conditions changes the methylation and expression patterns of imprinted genes [81,88]. In laboratory animals, 5–10% of non-manipulated
embryos undergoes abnormal methylation reprogramming and
fails to develop. However, embryos derived from some kinds of
manipulation, for example, superovulation and in vitro culture,
undoubtedly present a higher rate of methylation and/or
imprinting abnormalities when compared to non-manipulated
embryos [89,90]. When nuclear transfer is considered, methylation
patterns are also abnormal and highly variable between individuals
[91,92].
Imprinted loci disruption has been observed in a number of
human developmental disorders and cancers [reviewed by [93]].
For example, a loss of imprinting (LOI) has been found in patients
with PWS (at a frequency of approximately 1%), patients with AS (at
a frequency of 3%), patients with BWS (50% of patients), and nearly
50% of the transient neonatal diabetes mellitus [reviewed by
[44,67]].
The observation that epigenetic abnormalities are present in
normal or manipulated pregnancies has made the animal model
suitable for a more profound study of these perturbations. In this
6. Transgenic strategies to study mammalian development
F.F. Bressan et al. / Placenta 30 (2009) 823–834
a DNA sequence flanked by homologous sequences into the gene to
be inactivated [113]. Reporter genes, as well as antibiotic resistance
genes, are introduced into the center of the target gene, causing
interference with expression and also allowing for the positive
selection of the transgene in the cell genome [114].
Homologous recombination of plasmid and DNA sequences is
obtained with a very low and variable efficiency rate [115]. Normally, it consists of the recombination of similar chromosome
sections derived from each parent [116]. Gene-targeting technologies exploit this characteristic by recombining transgenes containing a disrupted gene with a similar DNA sequence, leading to
targeted gene disruption.
Successfully modified embryonic stem cells are injected in preimplantation blastocoels, contributing to the tissues of the developing animal, including the germline [117,118]. Embryonic and
adult tissues are composed of transgenic and non-transgenic cells
called chimeras. Once these embryonic stem cells are integrated
into germ cells, the newly inserted gene alteration may be passed
on to the next generations. As a result, the chimeras produced are
able to generate mouse strains that are heterozygous for the altered
genes, and, most importantly, homozygous offspring can be
obtained by planned matings [reviewed by [94]].
7. Developmental studies based on knockout models
Transgenic approaches in mice have provided reliable means of
investigating complex biological phenomena or diseases by
allowing gene products to be expressed in a controlled manner in
a whole organism where the majority of the genes have a human
counterpart [119,120].
Indeed, an International Mouse Knockout Consortium
composed of four groups, the Knockout Mouse Project (KOMP,
http://knockoutmouse.or), the European Conditional Mouse
Mutagenesis Program (EUCOMM, http://www.eucomm.org), the
North American Conditional Mouse Mutagenesis Program (NorCOMM, http://norcomm.phenogenomics.ca/index.htm), and the
Texas Institute for Genomic Medicine (TIGM, http://tigm.org), was
created in 2007 to obtain a mutation of all protein-encoding genes
in the mouse using a combination of gene-targeting and genetrapping strategies [96,121].
Regarding developmental process, mouse mutants have been
created for the broad study of gene expression and developmental
interactions not only throughout the peri-implantation and gestation periods [reviewed by [1,9]] but also for different stages of
reproduction [reviewed by [107,122]].
KO models have been used for more than a decade to investigate gene function, including the role of certain genes for epigenetic patterning and embryogenesis. Trasler and collaborators in
1996 [123] showed that DNA methyltransferase (Dnmt/) KO
embryos failed to develop past the 25-somite stage and were
developmentally delayed and asynchronous. The authors
concluded that DNA methylation is vital for embryo development.
829
Five main mammalian DNA methyltransferases (Dnmt) have been
characterized and are related to the establishment and maintenance of genomic imprinting: Dnmt1, Dnmt1o, Dnmt3a, Dnmt3b
and Dnmt3L [124]. Dnmt1 and the oocyte isoform Dnmt1o are
responsible for the maintenance of the imprinted methylation
patterns [125,126], Dnmt3a and Dnmt3b are required for de novo
methylation and are essential for paternal and maternal methylation imprints during germline development [127]. Most recent
studies have shown that Dnmt3-like (Dnmt3L) cooperates with
Dnmt3a and is necessary for the establishment of genomic
imprinting during gametogenesis [128–130]. By constructing KO
mice models, it was possible to show that these methyltransferases are indispensable for embryogenesis, as summarized
in Table 3.
Similar to DNA methylation, histone modifications, mainly
acetylation and methylation, also influence gene expression
[131,132]. In contrast to embryo formation, placentation seems to
be more dependent on repressive histone methylation than DNA
methylation, as stated earlier. Some imprints in extraembryonic
tissues directly correlated with histone H3 repressive methylation
but not with DNA methylation [133,134]. In placenta, several genes
maintain imprinting status in the absence of Dnmt1 [135,136].
These genes probably have their DNA-methylated allele enriched
with histone H3-lysine-9 methylation, together with other histone
lysine methylation. Using KO models, the histone methyltransferase (HMT) G9a was shown to contribute to the allelic repression of
genes that are imprinted only in the trophoblast. The dependence
of histone post-translational modification in the parental originspecific expression probably prevents imprint erasure during the
genome-wide demethylation wave that occurs after fertilization
[29,136].
KO studies of other HMTs or histone deacetylases (HDACs) have
shown that deletions of its encoding genes (i.e., HMTs Eset and G9a
and Polycomb-group genes Ezh2 and Suz12) lead to embryonic
lethality [131,137–139]. The mechanisms by which histone modifiers regulate the maintenance of differentially allelic chromatin
organization in imprints require further investigation.
From more than 70 imprinted genes in which expression was
already reported in developmental stages, roughly half have been
analyzed through KO studies, which are summarized in Table 2. The
phenotypes observed in the KOs ranged from increased embryonic
or post-natal lethality (i.e., Gnas, Peg10, Klf4, Ascl2, Tapa1/CD81) to
no obvious phenotypes (i.e., Nesp, Peg12/Frat3, Slc22a2). Most
phenotypes evaluated by KO experiments confirm the preferential
allelic gene expression and its importance for fetoplacental growth.
For example, Table 2 shows that the deletion of the paternally
expressed genes Peg10, Mest/Peg1, Peg3, Igf2, Dlk1, Gtl2, Rtl1 and
others suppresses growth, whereas the deletion of the maternally
expressed genes H19, Grb10, Igf2r and others increases fetoplacental growth. Some mutations, although apparently unrelated to
nutrition allocation and fetal growth, are essential for fetal development, i.e., the deletion of Ube3a, Sgce, Ppp1r9a and Pwcr1, among
Table 3
DNA and histone methyltransferases knockout consequences.
Methyltransferases
Knockout consequences
References
Dnmt1
Dnmt1o
Dnmt3a
Dnmt3b
Dnmt3a and Dnmt3b
Dnmt3L
Embryonic extensive demethylation
Embryos from Dnmt1o/ females lose half of their imprints during one cell cycle
Apparently normal at birth, increased lethality at about 4 weeks of age, presenting runted phenotype
Embryonic lethality probably due to multiple developmental defects
Impaired de novo methylation. Embryonic lethality before 11.5 dpc
Null mutations reveal disruption of maternal methylation imprints. Heterozygous
progeny of homozygous females fail to develop beyond 10.5 dpc
due to abnormal development of extraembryonic structures
Decrease in H3-K9 methylation in placenta, embryonic lethality before/at 10 dpc
[240]
[126]
[127,241]
[241]
[241,242]
[128–130]
HMT G9a
[29,137]
830
F.F. Bressan et al. / Placenta 30 (2009) 823–834
others, and mainly result in impairments related to the nervous
system during the post-natal period.
Interestingly, the deletion of Peg10, a paternally expressed gene,
as well as the deletion of Ascl2, a maternally expressed gene, both
leads to embryonic lethality due to placental defects. Fetal growth
and placentation are now seen as complex processes dependent on
very particular gene expression networks. By generating animals
lacking a specific gene, it was possible to evaluate a variety of
reproductive parameters in controlled experiments, turning
transgenesis into an extremely valuable tool for imprinted gene
expression studies.
8. Conclusions and perspectives
In 2007, the Nobel Prize in physiology or medicine was awarded
Drs. Mario Capechi, Martin Evans and Oliver Smithies for their work
on genetic modifications in mice using embryonic stem cells. Great
progress in several fields of basic and medical science was made
possible with the use of animals harboring genetic modifications.
Undoubtedly, this technology has greatly contributed to the
understanding of the mechanisms that regulate genomic
imprinting and development in mammals.
For a long time, the KO approach has been the method of choice
in placentation and early development studies, allowing for the
evaluation of specific phenotypes in vivo throughout gestation.
Although this technique is well established in mice, its technical
unavailability in other animal species is a considerable drawback.
Moreover, animals other than the mice have increasingly been
accepted as research models because they may be better correlated
with human characteristics such as birth weight, organ
morphology or genome similarity (i.e., ewe, swine or primate
models). We believe that in the near future, epigenome interferences, i.e., targeted epimutations in numerous animal models, may
allow the ‘‘knockout’’ technique to become the basis for several
other new and valuable techniques in science.
By reviewing the importance of genomic imprinting in early
development in mammals and the genes involved, we have
emphasized the role of imprinted genes in successful placental and
fetal development. Moreover, we have highlighted the regulation of
some important genes, which may turn into future targets of
genetic therapies.
Because the acquisition and evolution of genomic imprinting are
among the most fundamental biological questions, further use of
gene transfer techniques to improve the understanding of this
process in mammals is warranted. In particular, gaining insight into
the regulation of epigenetic mechanisms during early development
would greatly contribute to the improvement of ARTs and their
outcomes.
References
[1] Rinkenberger JL, Cross JC, Werb Z. Molecular genetics of implantation in the
mouse. Dev Genet 1997;21:6–20.
[2] Dantzer V. Epitheliochorial placentation. In: Encyclopedia of reproduction.
Academic Press; 1999.
[3] Miozzo M, Simoni G. The role of imprinted genes in fetal growth. Biol
Neonate 2002;81:217–28.
[4] Fowden AL, Sibley C, Reik W, Constancia M. Imprinted genes, placental
development and fetal growth. Horm Res 2006;65(Suppl. 3):50–8.
[5] Coan PM, Burton GJ, Ferguson-Smith AC. Imprinted genes in the placenta –
a review. Placenta 2005;26(Suppl. A):S10–20.
[6] Wagschal A, Feil R. Genomic imprinting in the placenta. Cytogenet Genome
Res 2006;113:90–8.
[7] Carter AM. Animal models of human placentation – a review. Placenta
2007;28(Suppl. A):S41–7.
[8] Cross JC. How to make a placenta: mechanisms of trophoblast cell differentiation in mice – a review. Placenta 2005;26(Suppl. A):S3–9.
[9] Rossant J, Cross JC. Placental development: lessons from mouse mutants. Nat
Rev Genet 2001;2:538–48.
[10] Hemberger M, Cross JC. Genes governing placental development. Trends
Endocrinol Metab 2001;12:162–8.
[11] Watson ED, Cross JC. Development of structures and transport functions in
the mouse placenta. Physiology (Bethesda) 2005;20:180–93.
[12] Nothias JY, Majumder S, Kaneko KJ, DePamphilis ML. Regulation of gene
expression at the beginning of mammalian development. J Biol Chem
1995;270:22077–80.
[13] Sood R, Zehnder JL, Druzin ML, Brown PO. Gene expression patterns in
human placenta. Proc Natl Acad Sci U S A 2006;103:5478–83.
[14] Nafee TM, Farrell WE, Carroll WD, Fryer AA, Ismail KM. Epigenetic control of
fetal gene expression. BJOG 2008;115:158–68.
[15] Ohgane J, Yagi S, Shiota K. Epigenetics: the DNA methylation profile of tissuedependent and differentially methylated regions in cells. Placenta
2008;29(Suppl. A):S29–35.
[16] Burton GJ, Kaufmann P, Huppertz B. Anatomy and genesis of the placenta. In:
Neills JD, editor. Knobil and Neill’s physiology of reproduction. Elsevier; 2006.
[17] Mann MR, Bartolomei MS. Epigenetic reprogramming in the mammalian
embryo: struggle of the clones. Genome Biol 2002;3. REVIEWS1003.
[18] Jones PA, Takai D. The role of DNA methylation in mammalian epigenetics.
Science 2001;293:1068–70.
[19] Delcuve GP, Rastegar M, Davie JR. Epigenetic control. J Cell Physiol
2009;219:243–50.
[20] Mayer W, Niveleau A, Walter J, Fundele R, Haaf T. Demethylation of the
zygotic paternal genome. Nature 2000;403:501–2.
[21] Tremblay KD, Saam JR, Ingram RS, Tilghman SM, Bartolomei MS. A paternalspecific methylation imprint marks the alleles of the mouse H19 gene. Nat
Genet 1995;9:407–13.
[22] Franklin GC, Adam GI, Ohlsson R. Genomic imprinting and mammalian
development. Placenta 1996;17:3–14.
[23] Paoloni-Giacobino A. Epigenetics in reproductive medicine. Pediatr Res
2007;61:51R–7R.
[24] Ferguson-Smith AC, Surani MA. Imprinting and the epigenetic asymmetry
between parental genomes. Science 2001;293:1086–9.
[25] Reik W, Constancia M, Fowden A, Anderson N, Dean W, Ferguson-Smith A,
et al. Regulation of supply and demand for maternal nutrients in mammals
by imprinted genes. J Physiol 2003;547:35–44.
[26] Constancia M, Angiolini E, Sandovici I, Smith P, Smith R, Kelsey G, et al.
Adaptation of nutrient supply to fetal demand in the mouse involves interaction between the Igf2 gene and placental transporter systems. Proc Natl
Acad Sci U S A 2005;102:19219–24.
[27] Branco MR, Oda M, Reik W. Safeguarding parental identity: Dnmt1 maintains
imprints during epigenetic reprogramming in early embryogenesis. Genes
Dev 2008;22:1567–71.
[28] Buiting K, Saitoh S, Gross S, Dittrich B, Schwartz S, Nicholls RD, et al.
Inherited microdeletions in the Angelman and Prader–Willi syndromes
define an imprinting centre on human chromosome 15. Nat Genet
1995;9:395–400.
[29] Wagschal A, Sutherland HG, Woodfine K, Henckel A, Chebli K, Schulz R, et al.
G9a histone methyltransferase contributes to imprinting in the mouse
placenta. Mol Cell Biol 2008;28:1104–13.
[30] Luedi PP, Dietrich FS, Weidman JR, Bosko JM, Jirtle RL, Hartemink AJ.
Computational and experimental identification of novel human imprinted
genes. Genome Res 2007;17:1723–30.
[31] Arnold DR, Lefebvre R, Smith LC. Characterization of the placenta specific
bovine mammalian achaete scute-like homologue 2 (Mash2) gene. Placenta
2006;27:1124–31.
[32] Zhang S, Kubota C, Yang L, Zhang Y, Page R, O’Neill M, et al. Genomic
imprinting of H19 in naturally reproduced and cloned cattle. Biol Reprod
2004;71:1540–4.
[33] Curchoe C, Zhang S, Bin Y, Zhang X, Yang L, Feng D, et al. Promoter-specific
expression of the imprinted IGF2 gene in cattle (Bos taurus). Biol Reprod
2005;73:1275–81.
[34] Dindot SV, Kent KC, Evers B, Loskutoff N, Womack J, Piedrahita JA. Conservation of genomic imprinting at the XIST, IGF2, and GTL2 loci in the bovine.
Mamm Genome 2004;15:966–74.
[35] Isles AR, Holland AJ. Imprinted genes and mother–offspring interactions.
Early Hum Dev 2005;81:73–7.
[36] Ferguson-Smith AC, Moore T, Detmar J, Lewis A, Hemberger M, Jammes H,
et al. Epigenetics and imprinting of the trophoblast – a workshop report.
Placenta 2006;27(Suppl. A):S122–6.
[37] Wood AJ, Oakey RJ. Genomic imprinting in mammals: emerging themes and
established theories. PLoS Genet 2006;2:e147.
[38] Dean W, Ferguson-Smith A. Genomic imprinting: mother maintains methylation marks. Curr Biol 2001;11:R527–30.
[39] Haig D, Graham C. Genomic imprinting and the strange case of the insulinlike growth factor II receptor. Cell 1991;64:1045–6.
[40] Moore T, Haig D. Genomic imprinting in mammalian development:
a parental tug-of-war. Trends Genet 1991;7:45–9.
[41] Barton SC, Surani MA, Norris ML. Role of paternal and maternal genomes in
mouse development. Nature 1984;311:374–6.
[42] Toppings M, Castro C, Mills PH, Reinhart B, Schatten G, Ahrens ET, et al.
Profound phenotypic variation among mice deficient in the maintenance of
genomic imprints. Hum Reprod 2008;23:807–18.
[43] Lucifero D, Mann MR, Bartolomei MS, Trasler JM. Gene-specific timing and
epigenetic memory in oocyte imprinting. Hum Mol Genet 2004;13:839–49.
F.F. Bressan et al. / Placenta 30 (2009) 823–834
[44] Horsthemke B, Ludwig M. Assisted reproduction: the epigenetic perspective.
Hum Reprod Update 2005;11:473–82.
[45] DeChiara TM, Robertson EJ, Efstratiadis A. Parental imprinting of the mouse
insulin-like growth factor II gene. Cell 1991;64:849–59.
[46] Constancia M, Hemberger M, Hughes J, Dean W, Ferguson-Smith A,
Fundele R, et al. Placental-specific IGF-II is a major modulator of placental
and fetal growth. Nature 2002;417:945–8.
[47] Kalscheuer VM, Mariman EC, Schepens MT, Rehder H, Ropers HH. The
insulin-like growth factor type-2 receptor gene is imprinted in the mouse but
not in humans. Nat Genet 1993;5:74–8.
[48] Killian JK, Nolan CM, Wylie AA, Li T, Vu TH, Hoffman AR, et al. Divergent
evolution in M6P/IGF2R imprinting from the Jurassic to the Quaternary. Hum
Mol Genet 2001;10:1721–8.
[49] Ludwig T, Eggenschwiler J, Fisher P, D’Ercole AJ, Davenport ML, Efstratiadis A.
Mouse mutants lacking the type 2 IGF receptor (IGF2R) are rescued from
perinatal lethality in Igf2 and Igf1r null backgrounds. Dev Biol 1996;177:
517–35.
[50] Coan PM, Fowden AL, Constancia M, Ferguson-Smith AC, Burton GJ, Sibley CP.
Disproportional effects of Igf2 knockout on placental morphology and
diffusional exchange characteristics in the mouse. J Physiol 2008;586:
5023–32.
[51] Srivastava M, Hsieh S, Grinberg A, Williams-Simons L, Huang SP, Pfeifer K.
H19 and Igf2 monoallelic expression is regulated in two distinct ways by
a shared cis acting regulatory region upstream of H19. Genes Dev
2000;14:1186–95.
[52] Delaval K, Feil R. Epigenetic regulation of mammalian genomic imprinting.
Curr Opin Genet Dev 2004;14:188–95.
[53] Thorvaldsen JL, Duran KL, Bartolomei MS. Deletion of the H19 differentially
methylated domain results in loss of imprinted expression of H19 and Igf2.
Genes Dev 1998;12:3693–702.
[54] Kono T, Obata Y, Wu Q, Niwa K, Ono Y, Yamamoto Y, et al. Birth of parthenogenetic mice that can develop to adulthood. Nature 2004;428:860–4.
[55] Obata Y, Kaneko-Ishino T, Koide T, Takai Y, Ueda T, Domeki I, et al.
Disruption of primary imprinting during oocyte growth leads to the
modified expression of imprinted genes during embryogenesis. Development 1998;125:1553–60.
[56] Moore T, Ball M. Kaguya, the first parthenogenetic mammal – engineering
triumph or lottery winner? Reproduction 2004;128:1–3.
[57] Sleutels F, Zwart R, Barlow DP. The non-coding Air RNA is required for
silencing autosomal imprinted genes. Nature 2002;415:810–3.
[58] Royo H, Cavaille J. Non-coding RNAs in imprinted gene clusters. Biol Cell
2008;100:149–66.
[59] Zhang Y, Qu L. Non-coding RNAs and the acquisition of genomic imprinting
in mammals. Sci China C Life Sci 2009;52:195–204.
[60] Edwards CA, Ferguson-Smith AC. Mechanisms regulating imprinted genes in
clusters. Curr Opin Cell Biol 2007;19:281–9.
[61] Peters J, Robson JE. Imprinted noncoding RNAs. Mamm Genome
2008;19:493–502.
[62] Mercer TR, Dinger ME, Mattick JS. Long non-coding RNAs: insights into
functions. Nat Rev Genet 2009;10:155–9.
[63] Ono R, Kobayashi S, Wagatsuma H, Aisaka K, Kohda T, Kaneko-Ishino T, et al.
A retrotransposon-derived gene, PEG10, is a novel imprinted gene located on
human chromosome 7q21. Genomics 2001;73:232–7.
[64] Ono R, Nakamura K, Inoue K, Naruse M, Usami T, Wakisaka-Saito N, et al.
Deletion of Peg10, an imprinted gene acquired from a retrotransposon,
causes early embryonic lethality. Nat Genet 2006;38:101–6.
[65] Ono R, Shiura H, Aburatani H, Kohda T, Kaneko-Ishino T, Ishino F. Identification of a large novel imprinted gene cluster on mouse proximal chromosome 6. Genome Res 2003;13:1696–705.
[66] Bliek J, Terhal P, van den Bogaard MJ, Maas S, Hamel B, Salieb-Beugelaar G,
et al. Hypomethylation of the H19 gene causes not only Silver–Russell
syndrome (SRS) but also isolated asymmetry or an SRS-like phenotype. Am J
Hum Genet 2006;78:604–14.
[67] Amor DJ, Halliday J. A review of known imprinting syndromes and their
association with assisted reproduction technologies. Hum Reprod
2008;23:2826–34.
[68] Paoloni-Giacobino A. Implications of reproductive technologies for birth and
developmental outcomes: imprinting defects and beyond. Expert Rev Mol
Med 2006;8:1–14.
[69] Capecchi MR. Altering the genome by homologous recombination. Science
1989;244:1288–92.
[70] Wilcox AJ, Baird DD, Weinberg CR. Time of implantation of the conceptus and
loss of pregnancy. N Engl J Med 1999;340:1796–9.
[71] Sapin V, Blanchon L, Serre AF, Lemery D, Dastugue B, Ward SJ. Use of
transgenic mice model for understanding the placentation: towards clinical
applications in human obstetrical pathologies? Transgenic Res 2001;10:
377–98.
[72] Cross JC, Werb Z, Fisher SJ. Implantation and the placenta: key pieces of the
development puzzle. Science 1994;266:1508–18.
[73] Ishiwata H, Katsuma S, Kizaki K, Patel OV, Nakano H, Takahashi T, et al.
Characterization of gene expression profiles in early bovine pregnancy using
a custom cDNA microarray. Mol Reprod Dev 2003;65:9–18.
[74] Andersen AN, Goossens V, Gianaroli L, Felberbaum R, Mouzon Jd, Nygren KG.
Assisted reproductive technology in Europe, 2003. Results generated from
European registers by ESHRE. Hum Reprod 2007;22:1513–25.
831
[75] Bertolini M, Anderson GB. The placenta as a contributor to production of
large calves. Theriogenology 2002;57:181–7.
[76] Constant F, Guillomot M, Heyman Y, Vignon X, Laigre P, Servely JL, et al. Large
offspring or large placenta syndrome? Morphometric analysis of late gestation bovine placentomes from somatic nuclear transfer pregnancies
complicated by hydrallantois. Biol Reprod 2006;75:122–30.
[77] Miglino MA, Pereira FT, Visintin JA, Garcia JM, Meirelles FV, Rumpf R, et al.
Placentation in cloned cattle: structure and microvascular architecture.
Theriogenology 2007;68:604–17.
[78] Arnold DR, Fortier AL, Lefebvre R, Miglino MA, Pfarrer C, Smith LC. Placental
insufficiencies in cloned animals – a workshop report. Placenta
2008;29(Suppl. A):S108–10.
[79] Young LE, Sinclair KD, Wilmut I. Large offspring syndrome in cattle and
sheep. Rev Reprod 1998;3:155–63.
[80] Ceelen M, Vermeiden JP. Health of human and livestock conceived by assisted
reproduction. Twin Res 2001;4:412–6.
[81] Young LE, Fernandes K, McEvoy TG, Butterwith SC, Gutierrez CG, Carolan C,
et al. Epigenetic change in IGF2R is associated with fetal overgrowth after
sheep embryo culture. Nat Genet 2001;27:153–4.
[82] Heyman Y, Chavatte-Palmer P, LeBourhis D, Camous S, Vignon X, Renard JP.
Frequency and occurrence of late-gestation losses from cattle cloned
embryos. Biol Reprod 2002;66:6–13.
[83] Hill JR, Burghardt RC, Jones K, Long CR, Looney CR, Shin T, et al. Evidence for
placental abnormality as the major cause of mortality in first-trimester
somatic cell cloned bovine fetuses. Biol Reprod 2000;63:1787–94.
[84] De Sousa PA, King T, Harkness L, Young LE, Walker SK, Wilmut I. Evaluation of
gestational deficiencies in cloned sheep fetuses and placentae. Biol Reprod
2001;65:23–30.
[85] Hill JR. Abnormal in utero development of cloned animals: implications for
human cloning. Differentiation 2002;69:174–8.
[86] Chavatte-Palmer P, Heyman Y, Richard C, Monget P, LeBourhis D, Kann G,
et al. Clinical, hormonal, and hematologic characteristics of bovine calves
derived from nuclei from somatic cells. Biol Reprod 2002;66:1596–603.
[87] Loi P, Clinton M, Vackova I, Fulka Jr J, Feil R, Palmieri C, et al. Placental
abnormalities associated with post-natal mortality in sheep somatic cell
clones. Theriogenology 2006;65:1110–21.
[88] Manipalviratn S, DeCherney A, Segars J. Imprinting disorders and assisted
reproductive technology. Fertil Steril 2009;91:305–15.
[89] Shi W, Haaf T. Aberrant methylation patterns at the two-cell stage as an
indicator of early developmental failure. Mol Reprod Dev 2002;63:
329–34.
[90] Fortier AL, Lopes FL, Darricarrere N, Martel J, Trasler JM. Superovulation alters
the expression of imprinted genes in the midgestation mouse placenta. Hum
Mol Genet 2008;17:1653–65.
[91] Kang YK, Koo DB, Park JS, Choi YH, Kim HN, Chang WK, et al. Typical
demethylation events in cloned pig embryos. Clues on species-specific
differences in epigenetic reprogramming of a cloned donor genome. J Biol
Chem 2001;276:39980–4.
[92] Ono Y, Shimozawa N, Muguruma K, Kimoto S, Hioki K, Tachibana M, et al.
Production of cloned mice from embryonic stem cells arrested at metaphase.
Reproduction 2001;122:731–6.
[93] Feinberg AP, Oshimura M, Barrett JC. Epigenetic mechanisms in human
disease. Cancer Res 2002;62:6784–7.
[94] Hacking DF. ‘Knock, and it shall be opened’: knocking out and knocking in to
reveal mechanisms of disease and novel therapies. Early Hum Dev
2008;84:821–7.
[95] Evans MJ, Carlton MB, Russ AP. Gene trapping and functional genomics.
Trends Genet 1997;13:370–4.
[96] Collins FS, Rossant J, Wurst W. A mouse for all reasons. Cell 2007;128:9–13.
[97] Georgiades P, Cox B, Gertsenstein M, Chawengsaksophak K, Rossant J.
Trophoblast-specific gene manipulation using lentivirus-based vectors. Biotechniques 2007;42: 317–18, 320, 322–5.
[98] Arroyo JA, Konno T, Khalili DC, Soares MJ. A simple in vivo approach to
investigate invasive trophoblast cells. Int J Dev Biol 2005;49:977–80.
[99] Jonkers J, van Amerongen R, van der Valk M, Robanus-Maandag E,
Molenaar M, Destree O, et al. In vivo analysis of Frat1 deficiency suggests
compensatory activity of Frat3. Mech Dev 1999;88:183–94.
[100] Hadjantonakis AK, Cox LL, Tam PP, Nagy A. An X-linked GFP transgene reveals
unexpected paternal X-chromosome activity in trophoblastic giant cells of
the mouse placenta. Genesis 2001;29:133–40.
[101] Wang J, Mager J, Chen Y, Schneider E, Cross JC, Nagy A, et al. Imprinted X
inactivation maintained by a mouse Polycomb group gene. Nat Genet
2001;28:371–5.
[102] Allemand I, Anguio J. Transgenic and knock-out models for studying DNA
repair. Biochimie 1995;77:826–32.
[103] Viney JL. Transgenic and gene knockout mice in cancer research. Cancer
Metastasis Rev 1995;14:77–90.
[104] Rees DA, Alcolado JC. Animal models of diabetes mellitus. Diabet Med
2005;22:359–70.
[105] Branchi I, Ricceri L. Transgenic and knock-out mouse pups: the growing need
for behavioral analysis. Genes Brain Behav 2002;1:135–41.
[106] Gregg AR. Mouse models and the role of nitric oxide in reproduction. Curr
Pharm Des 2003;9:391–8.
[107] Roy A, Matzuk MM. Deconstructing mammalian reproduction: using
knockouts to define fertility pathways. Reproduction 2006;131:207–19.
832
F.F. Bressan et al. / Placenta 30 (2009) 823–834
[108] Malassine A, Frendo JL, Evain-Brion D. A comparison of placental development and endocrine functions between the human and mouse model. Hum
Reprod Update 2003;9:531–9.
[109] Schulz R, Woodfine K, Menheniott TR, Bourc’his D, Bestor T, Oakey RJ.
WAMIDEX: a web atlas of murine genomic imprinting and differential
expression. Epigenetics 2008;3:89–96.
[110] Church DM, Goodstadt L, Hillier LW, Zody MC, Goldstein S, She X, et al.
Lineage-specific biology revealed by a finished genome assembly of the
mouse. PLoS Biol 2009;7. e1000112.
[111] Evans MJ, Kaufman MH. Establishment in culture of pluripotential cells from
mouse embryos. Nature 1981;292:154–6.
[112] Martin GR. Isolation of a pluripotent cell line from early mouse embryos
cultured in medium conditioned by teratocarcinoma stem cells. Proc Natl
Acad Sci U S A 1981;78:7634–8.
[113] Galli-Taliadoros LA, Sedgwick JD, Wood SA, Korner H. Gene knock-out technology: a methodological overview for the interested novice. J Immunol
Methods 1995;181:1–15.
[114] Habermann FA, Wuensch A, Sinowatz F, Wolf E. Reporter genes for
embryogenesis research in livestock species. Theriogenology 2007;68(Suppl.
1):S116–24.
[115] te Riele H, Maandag ER, Berns A. Highly efficient gene targeting in embryonic
stem cells through homologous recombination with isogenic DNA constructs.
Proc Natl Acad Sci U S A 1992;89:5128–32.
[116] Shibata T, Nishinaka T, Mikawa T, Aihara H, Kurumizaka H, Yokoyama S, et al.
Homologous genetic recombination as an intrinsic dynamic property of
a DNA structure induced by RecA/Rad51-family proteins: a possible advantage
of DNA over RNA as genomic material. Proc Natl Acad Sci U S A
2001;98:8425–32.
[117] Gossler A, Doetschman T, Korn R, Serfling E, Kemler R. Transgenesis by means
of blastocyst-derived embryonic stem cell lines. Proc Natl Acad Sci U S A
1986;83:9065–9.
[118] Robertson E, Bradley A, Kuehn M, Evans M. Germ-line transmission of genes
introduced into cultured pluripotential cells by retroviral vector. Nature
1986;323:445–8.
[119] Misra RP, Duncan SA. Gene targeting in the mouse: advances in introduction
of transgenes into the genome by homologous recombination. Endocrine
2002;19:229–38.
[120] Sands AT. The master mammal. Nat Biotechnol 2003;21:31–2.
[121] Collins FS, Finnell RH, Rossant J, Wurst W. A new partner for the international
knockout mouse consortium. Cell 2007;129:235.
[122] Matzuk MM, Lamb DJ. Genetic dissection of mammalian fertility pathways.
Nat Cell Biol 2002;4(Suppl.):s41–9.
[123] Trasler JM, Trasler DG, Bestor TH, Li E, Ghibu F. DNA methyltransferase in
normal and Dnmtn/Dnmtn mouse embryos. Dev Dyn 1996;206:239–47.
[124] Chen T, Li E. Structure and function of eukaryotic DNA methyltransferases.
Curr Top Dev Biol 2004;60:55–89.
[125] Leonhardt H, Page AW, Weier HU, Bestor TH. A targeting sequence directs
DNA methyltransferase to sites of DNA replication in mammalian nuclei. Cell
1992;71:865–73.
[126] Howell CY, Bestor TH, Ding F, Latham KE, Mertineit C, Trasler JM, et al.
Genomic imprinting disrupted by a maternal effect mutation in the Dnmt1
gene. Cell 2001;104:829–38.
[127] Kaneda M, Okano M, Hata K, Sado T, Tsujimoto N, Li E, et al. Essential role for
de novo DNA methyltransferase Dnmt3a in paternal and maternal
imprinting. Nature 2004;429:900–3.
[128] Hata K, Okano M, Lei H, Li E. Dnmt3L cooperates with the Dnmt3 family of de
novo DNA methyltransferases to establish maternal imprints in mice.
Development 2002;129:1983–93.
[129] Arima T, Hata K, Tanaka S, Kusumi M, Li E, Kato K, et al. Loss of the maternal
imprint in Dnmt3Lmat/ mice leads to a differentiation defect in the
extraembryonic tissue. Dev Biol 2006;297:361–73.
[130] Bourc’his D, Xu GL, Lin CS, Bollman B, Bestor TH. Dnmt3L and the establishment of maternal genomic imprints. Science 2001;294:2536–9.
[131] Dodge JE, Kang YK, Beppu H, Lei H, Li E. Histone H3-K9 methyltransferase ESET is essential for early development. Mol Cell Biol
2004;24:2478–86.
[132] Haberland M, Montgomery RL, Olson EN. The many roles of histone deacetylases in development and physiology: implications for disease and therapy.
Nat Rev Genet 2009;10:32–42.
[133] Lewis A, Mitsuya K, Umlauf D, Smith P, Dean W, Walter J, et al. Imprinting on
distal chromosome 7 in the placenta involves repressive histone methylation
independent of DNA methylation. Nat Genet 2004;36:1291–5.
[134] Umlauf D, Goto Y, Cao R, Cerqueira F, Wagschal A, Zhang Y, et al. Imprinting
along the Kcnq1 domain on mouse chromosome 7 involves repressive
histone methylation and recruitment of Polycomb group complexes. Nat
Genet 2004;36:1296–300.
[135] Caspary T, Cleary MA, Baker CC, Guan XJ, Tilghman SM. Multiple mechanisms
regulate imprinting of the mouse distal chromosome 7 gene cluster. Mol Cell
Biol 1998;18:3466–74.
[136] Tanaka M, Puchyr M, Gertsenstein M, Harpal K, Jaenisch R, Rossant J, et al.
Parental origin-specific expression of Mash2 is established at the time of
implantation with its imprinting mechanism highly resistant to genomewide demethylation. Mech Dev 1999;87:129–42.
[137] Tachibana M, Sugimoto K, Nozaki M, Ueda J, Ohta T, Ohki M, et al. G9a histone
methyltransferase plays a dominant role in euchromatic histone H3 lysine 9
[138]
[139]
[140]
[141]
[142]
[143]
[144]
[145]
[146]
[147]
[148]
[149]
[150]
[151]
[152]
[153]
[154]
[155]
[156]
[157]
[158]
[159]
[160]
[161]
[162]
[163]
methylation and is essential for early embryogenesis. Genes Dev
2002;16:1779–91.
O’Carroll D, Erhardt S, Pagani M, Barton SC, Surani MA, Jenuwein T. The
Polycomb-group gene Ezh2 is required for early mouse development. Mol
Cell Biol 2001;21:4330–6.
Lagger G, O’Carroll D, Rembold M, Khier H, Tischler J, Weitzer G, et al.
Essential function of histone deacetylase 1 in proliferation control and CDK
inhibitor repression. EMBO J 2002;21:2672–81.
Sandell LL, Guan XJ, Ingram R, Tilghman SM. Gatm, a creatine synthesis
enzyme, is imprinted in mouse placenta. Proc Natl Acad Sci U S A
2003;100:4622–7.
Evans HK, Wylie AA, Murphy SK, Jirtle RL. The neuronatin gene resides in
a ‘‘micro-imprinted’’ domain on human chromosome 20q11.2. Genomics
2001;77:99–104.
Peters J, Wroe SF, Wells CA, Miller HJ, Bodle D, Beechey CV, et al. A cluster of
oppositely imprinted transcripts at the Gnas locus in the distal imprinting
region of mouse chromosome 2. Proc Natl Acad Sci U S A 1999;96:3830–5.
Ball ST, Williamson CM, Hayes C, Hacker T, Peters J. The spatial and temporal
expression pattern of Nesp and its antisense Nespas, in mid-gestation mouse
embryos. Mech Dev 2001;100:79–81.
Wroe SF, Kelsey G, Skinner JA, Bodle D, Ball ST, Beechey CV, et al. An
imprinted transcript, antisense to Nesp, adds complexity to the cluster of
imprinted genes at the mouse Gnas locus. Proc Natl Acad Sci U S A
2000;97:3342–6.
Yu S, Yu D, Lee E, Eckhaus M, Lee R, Corria Z, et al. Variable and tissue-specific
hormone resistance in heterotrimeric Gs protein alpha-subunit (Gsalpha)
knockout mice is due to tissue-specific imprinting of the Gsalpha gene. Proc
Natl Acad Sci U S A 1998;95:8715–20.
Plagge A, Gordon E, Dean W, Boiani R, Cinti S, Peters J, et al. The imprinted
signaling protein XL alpha s is required for postnatal adaptation to feeding.
Nat Genet 2004;36:818–26.
Wood AJ, Bourc’his D, Bestor TH, Oakey RJ. Allele-specific demethylation at
an imprinted mammalian promoter. Nucleic Acids Res 2007;35:7031–9.
Kuzmin A, Han Z, Golding MC, Mann MR, Latham KE, Varmuza S. The PcG
gene Sfmbt2 is paternally expressed in extraembryonic tissues. Gene
Expression Patterns 2008;8:107–16.
Hoshiya H, Meguro M, Kashiwagi A, Okita C, Oshimura M. Calcr, a brainspecific imprinted mouse calcitonin receptor gene in the imprinted cluster of
the proximal region of chromosome 6. J Hum Genet 2003;48:208–11.
Lee YJ, Park CW, Hahn Y, Park J, Lee J, Yun JH, et al. Mit1/Lb9 and Copg2, new
members of mouse imprinted genes closely linked to Peg1/Mest(1). FEBS Lett
2000;472:230–4.
Piras G, El Kharroubi A, Kozlov S, Escalante-Alcalde D, Hernandez L,
Copeland NG, et al. Zac1 (Lot1), a potential tumor suppressor gene, and the
gene for epsilon-sarcoglycan are maternally imprinted genes: identification
by a subtractive screen of novel uniparental fibroblast lines. Mol Cell Biol
2000;20:3308–15.
Monk D, Wagschal A, Arnaud P, Muller PS, Parker-Katiraee L, Bourc’his D,
et al. Comparative analysis of human chromosome 7q21 and mouse proximal
chromosome 6 reveals a placental-specific imprinted gene, TFPI2/Tfpi2,
which requires EHMT2 and EED for allelic-silencing. Genome Res
2008;18:1270–81.
Mizuno Y, Sotomaru Y, Katsuzawa Y, Kono T, Meguro M, Oshimura M, et al.
Asb4, Ata3, and Dcn are novel imprinted genes identified by high-throughput
screening using RIKEN cDNA microarray. Biochem Biophys Res Commun
2002;290:1499–505.
Reule M, Krause R, Hemberger M, Fundele R. Analysis of Peg1/Mest
imprinting in the mouse. Dev Genes Evol 1998;208:161–3.
Mayer W, Hemberger M, Frank HG, Grummer R, Winterhager E, Kaufmann P,
et al. Expression of the imprinted genes MEST/Mest in human and murine
placenta suggests a role in angiogenesis. Dev Dyn 2000;217:1–10.
Parker-Katiraee L, Carson AR, Yamada T, Arnaud P, Feil R, Abu-Amero SN, et al.
Identification of the imprinted KLF14 transcription factor undergoing
human-specific accelerated evolution. PLoS Genet 2007;3. e65.
Smith RJ, Dean W, Konfortova G, Kelsey G. Identification of novel imprinted
genes in a genome-wide screen for maternal methylation. Genome Res
2003;13:558–69.
Kim J, Bergmann A, Wehri E, Lu X, Stubbs L. Imprinting and evolution of two
Kruppel-type zinc-finger genes, ZIM3 and ZNF264, located in the PEG3/USP29
imprinted domain. Genomics 2001;77:91–8.
Lewis A, Green K, Dawson C, Redrup L, Huynh KD, Lee JT, et al. Epigenetic
dynamics of the Kcnq1 imprinted domain in the early embryo. Development
2006;133:4203–10.
Kim J, Bergmann A, Lucas S, Stone R, Stubbs L. Lineage-specific imprinting
and evolution of the zinc-finger gene ZIM2. Genomics 2004;84:47–58.
Kim J, Lu X, Stubbs L. Zim1, a maternally expressed mouse Kruppel-type
zinc-finger gene located in proximal chromosome 7. Hum Mol Genet 1999;8:
847–54.
Li LL, Szeto IY, Cattanach BM, Ishino F, Surani MA. Organization and parentof-origin-specific methylation of imprinted Peg3 gene on mouse proximal
chromosome 7. Genomics 2000;63:333–40.
Kim J, Noskov VN, Lu X, Bergmann A, Ren X, Warth T, et al. Discovery of
a novel, paternally expressed ubiquitin-specific processing protease gene
through comparative analysis of an imprinted region of mouse chromosome
7 and human chromosome 19q13.4. Genome Res 2000;10:1138–47.
F.F. Bressan et al. / Placenta 30 (2009) 823–834
[164] Buettner VL, Walker AM, Singer-Sam J. Novel paternally expressed intergenic
transcripts at the mouse Prader–Willi/Angelman syndrome locus. Mamm
Genome 2005;16:219–27.
[165] de los Santos T, Schweizer J, Rees CA, Francke U. Small evolutionarily
conserved RNA, resembling C/D box small nucleolar RNA, is transcribed from
PWCR1, a novel imprinted gene in the Prader–Willi deletion region, which Is
highly expressed in brain. Am J Hum Genet 2000;67:1067–82.
[166] Gray TA, Smithwick MJ, Schaldach MA, Martone DL, Graves JA, McCarrey JR,
et al. Concerted regulation and molecular evolution of the duplicated SNRPB0 /
B and SNRPN loci. Nucleic Acids Res 1999;27:4577–84.
[167] Barr JA, Jones J, Glenister PH, Cattanach BM. Ubiquitous expression and
imprinting of Snrpn in the mouse. Mamm Genome 1995;6:405–7.
[168] Szabo PE, Mann JR. Biallelic expression of imprinted genes in the mouse
germ line: implications for erasure, establishment, and mechanisms of
genomic imprinting. Genes Dev 1995;9:1857–68.
[169] Uetsuki T, Takagi K, Sugiura H, Yoshikawa K. Structure and expression of the
mouse necdin gene. Identification of a postmitotic neuron-restrictive core
promoter. J Biol Chem 1996;271:918–24.
[170] Boccaccio I, Glatt-Deeley H, Watrin F, Roeckel N, Lalande M, Muscatelli F.
The human MAGEL2 gene and its mouse homologue are paternally
expressed and mapped to the Prader–Willi region. Hum Mol Genet
1999;8:2497–505.
[171] Hershko A, Razin A, Shemer R. Imprinted methylation and its effect on
expression of the mouse Zfp127 gene. Gene 1999;234:323–7.
[172] Rivera RM, Stein P, Weaver JR, Mager J, Schultz RM, Bartolomei MS. Manipulations of mouse embryos prior to implantation result in aberrant expression of imprinted genes on day 9.5 of development. Hum Mol Genet
2008;17:1–14.
[173] Jong MT, Gray TA, Ji Y, Glenn CC, Saitoh S, Driscoll DJ, et al. A novel imprinted
gene, encoding a RING zinc-finger protein, and overlapping antisense transcript in the Prader–Willi syndrome critical region. Hum Mol Genet
1999;8:783–93.
[174] Kobayashi S, Kohda T, Ichikawa H, Ogura A, Ohki M, Kaneko-Ishino T, et al.
Paternal expression of a novel imprinted gene, Peg12/Frat3, in the mouse 7C
region homologous to the Prader–Willi syndrome region. Biochem Biophys
Res Commun 2002;290:403–8.
[175] Choi JD, Underkoffler LA, Wood AJ, Collins JN, Williams PT, Golden JA, et al. A
novel variant of Inpp5f is imprinted in brain, and its expression is correlated
with differential methylation of an internal CpG island. Mol Cell Biol
2005;25:5514–22.
[176] Bartolomei MS, Zemel S, Tilghman SM. Parental imprinting of the mouse H19
gene. Nature 1991;351:153–5.
[177] Walsh C, Glaser A, Fundele R, Ferguson-Smith A, Barton S, Surani MA, et al.
The non-viability of uniparental mouse conceptuses correlates with the loss
of the products of imprinted genes. Mech Dev 1994;46:55–62.
[178] Deltour L, Montagutelli X, Guenet JL, Jami J, Paldi A. Tissue- and developmental stage-specific imprinting of the mouse proinsulin gene, Ins2. Dev Biol
1995;168:686–8.
[179] Duvillie B, Bucchini D, Tang T, Jami J, Paldi A. Imprinting at the mouse Ins2
locus: evidence for cis- and trans-allelic interactions. Genomics 1998;47:52–
7.
[180] Paulsen M, El-Maarri O, Engemann S, Strodicke M, Franck O, Davies K, et al.
Sequence conservation and variability of imprinting in the Beckwith–Wiedemann syndrome gene cluster in human and mouse. Hum Mol Genet
2000;9:1829–41.
[181] Engemann S, Strodicke M, Paulsen M, Franck O, Reinhardt R, Lane N, et al.
Sequence and functional comparison in the Beckwith–Wiedemann region:
implications for a novel imprinting centre and extended imprinting. Hum
Mol Genet 2000;9:2691–706.
[182] Lee MH, Reynisdottir I, Massague J. Cloning of p57KIP2, a cyclin-dependent
kinase inhibitor with unique domain structure and tissue distribution. Genes
Dev 1995;9:639–49.
[183] Dao D, Frank D, Qian N, O’Keefe D, Vosatka RJ, Walsh CP, et al. IMPT1, an
imprinted gene similar to polyspecific transporter and multi-drug resistance
genes. Hum Mol Genet 1998;7:597–608.
[184] Morisaki H, Hatada I, Morisaki T, Mukai T. A novel gene, ITM, located between
p57KIP2 and IPL, is imprinted in mice. DNA Res 1998;5:235–40.
[185] Qian N, Frank D, O’Keefe D, Dao D, Zhao L, Yuan L, et al. The IPL gene on
chromosome 11p15.5 is imprinted in humans and mice and is similar to
TDAG51, implicated in Fas expression and apoptosis. Hum Mol Genet
1997;6:2021–9.
[186] Frank D, Fortino W, Clark L, Musalo R, Wang W, Saxena A, et al. Placental
overgrowth in mice lacking the imprinted gene Ipl. Proc Natl Acad Sci U S A
2002;99:7490–5.
[187] Paulsen M, Davies KR, Bowden LM, Villar AJ, Franck O, Fuermann M, et al.
Syntenic organization of the mouse distal chromosome 7 imprinting cluster
and the Beckwith–Wiedemann syndrome region in chromosome 11p15.5.
Hum Mol Genet 1998;7:1149–59.
[188] Clark L, Wei M, Cattoretti G, Mendelsohn C, Tycko B. The Tnfrh1 (Tnfrsf23)
gene is weakly imprinted in several organs and expressed at the trophoblast–
decidua interface. BMC Genet 2002;3:11.
[189] Higashimoto K, Soejima H, Yatsuki H, Joh K, Uchiyama M, Obata Y, et al.
Characterization and imprinting status of OBPH1/Obph1 gene: implications
for an extended imprinting domain in human and mouse. Genomics
2002;80:575–84.
833
[190] Menheniott TR, Woodfine K, Schulz R, Wood AJ, Monk D, Giraud AS, et al.
Genomic imprinting of Dopa decarboxylase in heart and reciprocal allelic
expression with neighboring Grb10. Mol Cell Biol 2008;28:386–96.
[191] Miyoshi N, Kuroiwa Y, Kohda T, Shitara H, Yonekawa H, Kawabe T, et al.
Identification of the Meg1/Grb10 imprinted gene on mouse proximal chromosome 11, a candidate for the Silver–Russell syndrome gene. Proc Natl Acad
Sci U S A 1998;95:1102–7.
[192] Hatada I, Mukai T. Genomic imprinting of p57KIP2, a cyclin-dependent
kinase inhibitor, in mouse. Nat Genet 1995;11:204–6.
[193] Tierling S, Dalbert S, Schoppenhorst S, Tsai CE, Oliger S, Ferguson-Smith AC,
et al. High-resolution map and imprinting analysis of the Gtl2–Dnchc1
domain on mouse chromosome 12. Genomics 2006;87:225–35.
[194] Kobayashi S, Wagatsuma H, Ono R, Ichikawa H, Yamazaki M, Tashiro H, et al.
Mouse Peg9/Dlk1 and human PEG9/DLK1 are paternally expressed imprinted
genes closely located to the maternally expressed imprinted genes: mouse
Meg3/Gtl2 and human MEG3. Genes Cells 2000;5:1029–37.
[195] Schmidt JV, Matteson PG, Jones BK, Guan XJ, Tilghman SM. The Dlk1 and
Gtl2 genes are linked and reciprocally imprinted. Genes Dev
2000;14:1997–2002.
[196] Sekita Y, Wagatsuma H, Nakamura K, Ono R, Kagami M, Wakisaka N, et al.
Role of retrotransposon-derived imprinted gene, Rtl1, in the feto–maternal
interface of mouse placenta. Nat Genet 2008;40:243–8.
[197] Yevtodiyenko A, Carr MS, Patel N, Schmidt JV. Analysis of candidate
imprinted genes linked to Dlk1–Gtl2 using a congenic mouse line. Mamm
Genome 2002;13:633–8.
[198] Davis E, Caiment F, Tordoir X, Cavaille J, Ferguson-Smith A, Cockett N, et al.
RNAi-mediated allelic trans-interaction at the imprinted Rtl1/Peg11 locus.
Curr Biol 2005;15:743–9.
[199] Kato MV, Ikawa Y, Hayashizaki Y, Shibata H. Paternal imprinting of mouse
serotonin receptor 2A gene Htr2 in embryonic eye: a conserved imprinting
regulation on the RB/Rb locus. Genomics 1998;47:146–8.
[200] Ruf N, Bahring S, Galetzka D, Pliushch G, Luft FC, Nurnberg P, et al. Sequencebased bioinformatic prediction and QUASEP identify genomic imprinting of
the KCNK9 potassium channel gene in mouse and human. Hum Mol Genet
2007;16:2591–9.
[201] Verhaagh S, Schweifer N, Barlow DP, Zwart R. Cloning of the mouse and
human solute carrier 22a3 (Slc22a3/SLC22A3) identifies a conserved cluster of
three organic cation transporters on mouse chromosome 17 and human
6q26–q27. Genomics 1999;55:209–18.
[202] Zwart R, Verhaagh S, de Jong J, Lyon M, Barlow DP. Genetic analysis of the
organic cation transporter genes Orct2/Slc22a2 and Orct3/Slc22a3 reduces the
critical region for the t haplotype mutant t(w73) to 200 kb. Mamm Genome
2001;12:734–40.
[203] Barlow DP, Stoger R, Herrmann BG, Saito K, Schweifer N. The mouse insulinlike growth factor type-2 receptor is imprinted and closely linked to the Tme
locus. Nature 1991;349:84–7.
[204] Coan PM, Conroy N, Burton GJ, Ferguson-Smith AC. Origin and characteristics
of glycogen cells in the developing murine placenta. Dev Dyn
2006;235:3280–94.
[205] Seidl CI, Stricker SH, Barlow DP. The imprinted Air ncRNA is an atypical
RNAPII transcript that evades splicing and escapes nuclear export. EMBO J
2006;25:3565–75.
[206] Plagge A, Isles AR, Gordon E, Humby T, Dean W, Gritsch S, et al. Imprinted
Nesp55 influences behavioral reactivity to novel environments. Mol Cell Biol
2005;25:3019–26.
[207] Yu S, Gavrilova O, Chen H, Lee R, Liu J, Pacak K, et al. Paternal versus maternal
transmission of a stimulatory G-protein alpha subunit knockout produces
opposite effects on energy metabolism. J Clin Invest 2000;105:615–23.
[208] Yokoi F, Dang MT, Mitsui S, Li Y. Exclusive paternal expression and novel
alternatively spliced variants of epsilon-sarcoglycan mRNA in mouse brain.
FEBS Lett 2005;579:4822–8.
[209] Wu LJ, Ren M, Wang H, Kim SS, Cao X, Zhuo M. Neurabin contributes to
hippocampal long-term potentiation and contextual fear memory. PLoS ONE
2008;3. e1407.
[210] Lefebvre L, Viville S, Barton SC, Ishino F, Keverne EB, Surani MA. Abnormal
maternal behaviour and growth retardation associated with loss of the
imprinted gene Mest. Nat Genet 1998;20:163–9.
[211] Segre JA, Bauer C, Fuchs E. Klf4 is a transcription factor required for establishing the barrier function of the skin. Nat Genet 1999;22:356–60.
[212] Mancini-Dinardo D, Steele SJ, Levorse JM, Ingram RS, Tilghman SM. Elongation of the Kcnq1ot1 transcript is required for genomic imprinting of
neighboring genes. Genes Dev 2006;20:1268–82.
[213] Li L, Keverne EB, Aparicio SA, Ishino F, Barton SC, Surani MA. Regulation of
maternal behavior and offspring growth by paternally expressed Peg3.
Science 1999;284:330–3.
[214] Jiang YH, Armstrong D, Albrecht U, Atkins CM, Noebels JL, Eichele G, et al.
Mutation of the Angelman ubiquitin ligase in mice causes increased cytoplasmic p53 and deficits of contextual learning and long-term potentiation.
Neuron 1998;21:799–811.
[215] Ding F, Li HH, Zhang S, Solomon NM, Camper SA, Cohen P, et al. SnoRNA,
Snord116 (Pwcr1/MBII-85) deletion causes growth deficiency and hyperphagia in mice. PLoS ONE 2008;3. e1709.
[216] Yang T, Adamson TE, Resnick JL, Leff S, Wevrick R, Francke U, et al. A mouse
model for Prader–Willi syndrome imprinting-centre mutations. Nat Genet
1998;19:25–31.
834
F.F. Bressan et al. / Placenta 30 (2009) 823–834
[217] Gerard M, Hernandez L, Wevrick R, Stewart CL. Disruption of the mouse
necdin gene results in early post-natal lethality. Nat Genet 1999;23:199–202.
[218] Pagliardini S, Ren J, Wevrick R, Greer JJ. Developmental abnormalities of
neuronal structure and function in prenatal mice lacking the Prader–Willi
syndrome gene necdin. Am J Pathol 2005;167:175–91.
[219] Bischof JM, Stewart CL, Wevrick R. Inactivation of the mouse Magel2 gene
results in growth abnormalities similar to Prader–Willi syndrome. Hum Mol
Genet 2007;16:2713–9.
[220] van Amerongen R, Nawijn M, Franca-Koh J, Zevenhoven J, van der Gulden H,
Jonkers J, et al. Frat is dispensable for canonical Wnt signaling in mammals.
Genes Dev 2005;19:425–30.
[221] Leighton PA, Ingram RS, Eggenschwiler J, Efstratiadis A, Tilghman SM.
Disruption of imprinting caused by deletion of the H19 gene region in mice.
Nature 1995;375:34–9.
[222] Leroux L, Desbois P, Lamotte L, Duvillie B, Cordonnier N, Jackerott M, et al.
Compensatory responses in mice carrying a null mutation for Ins1 or Ins2.
Diabetes 2001;50(Suppl. 1):S150–3.
[223] Duvillie B, Cordonnier N, Deltour L, Dandoy-Dron F, Itier JM, Monthioux E,
et al. Phenotypic alterations in insulin-deficient mutant mice. Proc Natl Acad
Sci U S A 1997;94:5137–40.
[224] Guillemot F, Nagy A, Auerbach A, Rossant J, Joyner AL. Essential role of Mash2 in extraembryonic development. Nature 1994;371:333–6.
[225] Rubinstein E, Ziyyat A, Prenant M, Wrobel E, Wolf JP, Levy S, et al. Reduced
fertility of female mice lacking CD81. Dev Biol 2006;290:351–8.
[226] Lee MP, Ravenel JD, Hu RJ, Lustig LR, Tomaselli G, Berger RD, et al. Targeted
disruption of the Kvlqt1 gene causes deafness and gastric hyperplasia in
mice. J Clin Invest 2000;106:1447–55.
[227] Zhang P, Wong C, Liu D, Finegold M, Harper JW, Elledge SJ. p21(CIP1) and
p57(KIP2) control muscle differentiation at the myogenin step. Genes Dev
1999;13:213–24.
[228] Takahashi K, Kobayashi T, Kanayama N. p57(Kip2) regulates the proper
development of labyrinthine and spongiotrophoblasts. Mol Hum Reprod
2000;6:1019–25.
[229] Takahashi K, Nakayama K. Mice lacking a CDK inhibitor, p57Kip2, exhibit
skeletal abnormalities and growth retardation. J Biochem 2000;127:73–83.
[230] Varrault A, Gueydan C, Delalbre A, Bellmann A, Houssami S, Aknin C, et al.
Zac1 regulates an imprinted gene network critically involved in the control of
embryonic growth. Dev Cell 2006;11:711–22.
[231] Danielson KG, Baribault H, Holmes DF, Graham H, Kadler KE, Iozzo RV. Targeted disruption of decorin leads to abnormal collagen fibril morphology and
skin fragility. J Cell Biol 1997;136:729–43.
[232] Iozzo RV, Moscatello DK, McQuillan DJ, Eichstetter I. Decorin is a biological
ligand for the epidermal growth factor receptor. J Biol Chem 1999;274:
4489–92.
[233] Charalambous M, Smith FM, Bennett WR, Crew TE, Mackenzie F, Ward A.
Disruption of the imprinted Grb10 gene leads to disproportionate overgrowth by an Igf2-independent mechanism. Proc Natl Acad Sci U S A
2003;100:8292–7.
[234] Moon YS, Smas CM, Lee K, Villena JA, Kim KH, Yun EJ, et al. Mice lacking
paternally expressed Pref-1/Dlk1 display growth retardation and accelerated
adiposity. Mol Cell Biol 2002;22:5585–92.
[235] Schuster-Gossler K, Simon-Chazottes D, Guenet JL, Zachgo J, Gossler A.
Gtl2lacZ, an insertional mutation on mouse chromosome 12 with parental
origin-dependent phenotype. Mamm Genome 1996;7:20–4.
[236] Angiolini E, Fowden A, Coan P, Sandovici I, Smith P, Dean W, et al. Regulation
of placental efficiency for nutrient transport by imprinted genes. Placenta
2006;27(Suppl. A):S98–102.
[237] Jonker JW, Wagenaar E, Van Eijl S, Schinkel AH. Deficiency in the
organic cation transporters 1 and 2 (Oct1/Oct2 [Slc22a1/Slc22a2]) in
mice abolishes renal secretion of organic cations. Mol Cell Biol
2003;23:7902–8.
[238] Wang ZQ, Fung MR, Barlow DP, Wagner EF. Regulation of embryonic growth
and lysosomal targeting by the imprinted Igf2/Mpr gene. Nature
1994;372:464–7.
[239] Wutz A, Theussl HC, Dausman J, Jaenisch R, Barlow DP, Wagner EF. Nonimprinted Igf2r expression decreases growth and rescues the Tme mutation
in mice. Development 2001;128:1881–7.
[240] Lei H, Oh SP, Okano M, Juttermann R, Goss KA, Jaenisch R, et al. De novo DNA
cytosine methyltransferase activities in mouse embryonic stem cells.
Development 1996;122:3195–205.
[241] Okano M, Bell DW, Haber DA, Li E. DNA methyltransferases Dnmt3a and
Dnmt3b are essential for de novo methylation and mammalian development.
Cell 1999;99:247–57.
[242] Chen T, Ueda Y, Dodge JE, Wang Z, Li E. Establishment and maintenance of
genomic methylation patterns in mouse embryonic stem cells by Dnmt3a
and Dnmt3b. Mol Cell Biol 2003;23:5594–605.